Making and Breaking of Lead Halide Perovskites - Accounts of

Jan 20, 2016 - All-Inorganic Perovskite Solar Cells. Jia Liang , Caixing Wang , Yanrong Wang , Zhaoran Xu , Zhipeng Lu , Yue Ma , Hongfei Zhu , Yi Hu ...
44 downloads 10 Views 6MB Size
Article pubs.acs.org/accounts

Making and Breaking of Lead Halide Perovskites Published as part of the Accounts of Chemical Research special issue “Lead Halide Perovskites for Solar Energy Conversion”. Joseph S. Manser,†,‡ Makhsud I. Saidaminov,∥ Jeffrey A. Christians,†,‡ Osman M. Bakr,∥ and Prashant V. Kamat*,†,‡,§ †

Radiation Laboratory, ‡Department of Chemical and Biomolecular Engineering, and §Department of Chemistry and Biochemistry, University of Notre Dame, Notre Dame, Indiana 46556, United States ∥ Division of Physical Sciences and Engineering, Solar and Photovoltaics Engineering Center, King Abdullah University of Science and Technology (KAUST), Thuwal 23955-6900, Kingdom of Saudi Arabia CONSPECTUS: A new front-runner has emerged in the field of next-generation photovoltaics. A unique class of materials, known as organic metal halide perovskites, bridges the gap between low-cost fabrication and exceptional device performance. These compounds can be processed at low temperature (typically in the range 80−150 °C) and readily self-assemble from the solution phase into high-quality semiconductor thin films. The low energetic barrier for crystal formation has mixed consequences. On one hand, it enables inexpensive processing and both optical and electronic tunability. The caveat, however, is that many as-formed lead halide perovskite thin films lack chemical and structural stability, undergoing rapid degradation in the presence of moisture or heat. To date, improvements in perovskite solar cell efficiency have resulted primarily from better control over thin film morphology, manipulation of the stoichiometry and chemistry of lead halide and alkylammonium halide precursors, and the choice of solvent treatment. Proper characterization and tuning of processing parameters can aid in rational optimization of perovskite devices. Likewise, gaining a comprehensive understanding of the degradation mechanism and identifying components of the perovskite structure that may be particularly susceptible to attack by moisture are vital to mitigate device degradation under operating conditions. This Account provides insight into the lifecycle of organic−inorganic lead halide perovskites, including (i) the nature of the precursor solution, (ii) formation of solid-state perovskite thin films and single crystals, and (iii) transformation of perovskites into hydrated phases upon exposure to moisture. In particular, spectroscopic and structural characterization techniques shed light on the thermally driven evolution of the perovskite structure. By tuning precursor stoichiometry and chemistry, and thus the lead halide charge-transfer complexes present in solution, crystallization kinetics can be tailored to yield improved thin film homogeneity. Because degradation of the as-formed perovskite film is in many ways analogous to its initial formation, the same suite of monitoring techniques reveals the moisture-induced transformation of low band gap methylammonium lead iodide (CH3NH3PbI3) to wide band gap hydrate compounds. The rate of degradation is increased upon exposure to light. Interestingly, the hydration process is reversible under certain conditions. This facile formation and subsequent chemical lability raises the question of whether CH3NH3PbI3 and its analogues are thermodynamically stable phases, thus posing a significant challenge to the development of transformative perovskite photovoltaics. Adequately addressing issues of structural and chemical stability under real-world operating conditions is paramount if perovskite solar cells are to make an impact beyond the benchtop. Expanding our fundamental knowledge of lead halide perovskite formation and degradation pathways can facilitate fabrication of stable, high-quality perovskite thin films for the next generation of photovoltaic and light emitting devices.



INTRODUCTION Oxide perovskites, with general ABX3 stoichiometry, have earned the distinction of “chemical chameleons” given that over 50% of all elements can be incorporated into at least one of the three lattice sites.1,2 Adding to the extensive variety of inorganic oxide perovskites, substitution of halide ions in place of oxygen in the X-site and incorporation of organic molecules into the A-site yields an entirely new series of compounds known as organic− © XXXX American Chemical Society

inorganic hybrid metal halide perovskites. Chemical interactions between organic and inorganic components, both in the precursor solution and in the solid state, influence formation, optoelectronic properties, and stability of hybrid perovskites. The underlying mechanisms of these interactions, which have Received: October 6, 2015

A

DOI: 10.1021/acs.accounts.5b00455 Acc. Chem. Res. XXXX, XXX, XXX−XXX

Article

Accounts of Chemical Research profound impact on material properties, have only recently come to light. In the solution phase, altering precursor stoichiometry and/or chemistry yields different lead halide charge transfer complexes that heavily influence crystallization kinetics and thin film quality.3−8 Formation of Lewis acid−base adducts between coordinating solvents such as dimethyl sulfoxide and solvated Pb has been utilized to produce record efficiency solar cells.9,10 Thus, proper control over precursor coordination chemistry as a means to improve perovskite solar cell performance should not be underestimated. Within the hybrid perovskite lattice, organic cations can undergo crystal phase-dependent reorientation.11 In contrast to inorganic perovskites, these molecular ions often exhibit permanent molecular dipoles (2.3 D for gas phase CH3NH3+).12 Their collective rotations within inorganic cuboctahedral interstitial sites can alter the local dielectric environment, potentially screening Coulombically bound electrons and holes as well as establishing local ferroelectric domains.13,14 Additionally, hydrogen bonding between the organic and inorganic components influences structural integrity and facilitates interactions with water and other polar molecules.12,15−17 The following sections provide a look at how the unique chemistry of lead halide perovskites enables efficient fabrication of thin films and single crystals with exceptional semiconducting properties, but also renders these materials vulnerable to degradation.

Figure 1. (a) Absorption spectra of 250 μM PbI2 in DMF (i) with increasing CH3NH3I concentration from 6 mM (ii) to 24 mM (v). Adapted from ref 4 with permission from The Royal Society of Chemistry.

introduction of chloride ions into the precursor mixture, which alters not only the solution chemistry but also the growth dynamics of the subsequent film.5 We still know relatively little about how interactions in solution influence the solid-state product. As we will discuss in the following section, advancing our understanding of the link between solution coordination chemistry and subsequent perovskite formation can aid in the rational and high-throughput fabrication of perovskite thin films with predictable and reproducible properties.



COORDINATION CHEMISTRY OF PEROVSKITE PRECURSORS The chemistry of lead halide complexes is well established.18 Transition metals with an s2 electron configuration (e.g., Ti4+, Sn2+, Pb2+, Sb3+) readily undergo complexation with halide ions. Lead complexes are commonly referred to as “plumbates” (e.g., triiodoplumbate (PbI3−)) and serve as precursors in solution. Solvents such as dimethylformamide (DMF) and dimethyl sulfoxide (DMSO) are commonly used in preparation of perovskite films and can donate a pair of electrons to form Lewis adducts with Pb2+. When dissolved in DMSO, PbI2 is colorless but becomes darker yellow upon addition of excess iodide ions. As coordinating solvent ligands are replaced by I− we observe formation of PbI3− and PbI42− complexes that exhibit lower energy charge-transfer absorption bands.4 These new transitions can be readily tracked and provide insight into complexation events in the precursor solution. Figure 1 shows absorption spectra of PbI2 recorded at different I− concentrations in DMF. The dependence of absorption on the I− concentration enables estimation of complexation constants (equilibria 1 and 2). PbI 2 + I− ← K1 → PbI3−



Perovskite Thin Film Formation

The path from solution-based precursors to solid-state perovskite dictates many of the physical and optoelectronic properties exhibited in the final film.19 Varying the stoichiometric ratio of CH3NH3I:PbI2 from 3:1 to 1:1 in the precursor solution yields thin films with dramatically different properties.3 The absorption spectra of CH3NH3PbI3 films deposited from 3:1 and 1:1 precursor solutions on mesoporous Al2O3 at different annealing times are shown in Figure 2a and b, respectively. Prior to annealing of the 3:1 film, the spectral features are comparable to the iodoplumbate complexes observed in solution. This suggests that under these conditions, the initial film is composed of iodoplumbates with various degrees of coordination. Annealing at 150 °C provides the requisite energy to drive off excess CH3NH3I and gradually transform the solid-state precursor (SSP) into perovskite. Use of stoichiometric solutions does not yield the same iodoplumbate fingerprint, suggesting a fundamentally different formation pathway that does not proceed through an SSP. These findings are complimented by transient absorption spectroscopy which explores excited state properties of the SSP and perovskite phases. Time-resolved difference absorption spectra of a CH3NH3PbI3 film prepared from a 3:1 precursor solution before and after annealing are given in Figure 2c and d, respectively. Both samples exhibit three distinct spectral features in the probe window from 400 to 760 nm: a high energy photobleach (PB1), photoinduced absorption (PA), and low energy photobleach (PB2). The sample before annealing exhibits characteristic bleaching peaks corresponding to transitions in the ground state iodoplumbate complex. In the fully annealed film, PB1 transforms into a broader peak characteristic of the CH3NH3PbI3 excited state spectrum.20 Evolution of spectral

(1)

PbI3− + I− ← K 2 → PbI4 2 −

FROM SOLUTION TO THE SOLID-STATE

(2) −1

The complexation constant for the two equilibria are 54 M for K1 and 6 M−1 for K2 in DMF.4 It is interesting to note that the complexation of Pb2+ with iodide ions is independent of the nature of the iodide counterion. Similar spectral features corresponding to lead halide complexes were also seen when CH3NH3I was replaced with KI. There are a multitude of different perovskite precursor formulations developed to-date that include mixtures of solvents, halide and molecular ions, and solvated metals. It is important to be cognizant of all the various interactions that can occur between the different components. A well-known example is the B

DOI: 10.1021/acs.accounts.5b00455 Acc. Chem. Res. XXXX, XXX, XXX−XXX

Article

Accounts of Chemical Research

Figure 2. Steady-state absorption spectra of CH3NH3PbI3 thin films on mp-Al2O3 prepared using (a) 3:1 and (b) 1:1 CH3NH3I:PbI2 (molar) precursor solutions at various annealing times. Inset: photographs of CH3NH3PbI3 films after annealing for the specified duration. (c,d) Time resolved difference absorption spectra recorded following 387 nm laser pulse excitation of CH3NH3PbI3 films prepared from a 3:1 precursor solution. The spectra were recorded (c) before and (d) after annealing for 20 min at 150 °C. Adapted with permission from ref 3. Copyright 2015 American Chemical Society.

features during annealing reveals how plumbate complexes transform into perovskite following incorporation of the organic cation into the lattice. Upon initial deposition of the 3:1 precursor solution, an SSP phase that does not correspond to CH3NH3PbI3 or the individual precursor solids is formed (Figure 3a). Compared to the (001) plane of PbI2 (12.7° 2θ),21 the lower angle reflection in the SSP (11° 2θ) indicates a larger lattice spacing in the intermediate structure. Expansion may result from incorporation of some combination of methylammonium cations and DMF molecules between the inorganic sheets. Recent work on perovskite crystallization mechanisms indicates formation of crystalline intermediates such as (CH3NH3+)2(PbI3−)2·DMF2 during solution processing.8 Annealing of the iodoplumbate SSP phase at 150 °C causes precursor reflections to gradually decrease along with a concomitant rise in tetragonal CH3NH3PbI3 [001] peaks. Stepwise solid-state transition from crystalline SSP corroborates the gradual evolution of perovskite optical properties with use of excess alkylammonium halide salt (vide supra), suggesting that under these conditions perovskite formation proceeds according to Ostwald’s rule of stages. Conversely, use of equimolar precursor solutions results in instantaneous crystallization upon removal of DMF (Figure 3a), causing spherulitic growth with large areas of the underlying substrate exposed (Figure 3b). The SSP phase that manifests with use of 3:1 solutions retards perovskite crystallization, yielding thin films with vastly improved surface coverage and homogeneity (Figure 3c).

Formation of a metastable SSP upon solution deposition opens unique possibilities for tailoring perovskite thin film growth. Use of different lead salts, whether they contain halides or spectator counterions as in Pb(NO3)2 or Pb(CH3CO2)2, provides crystalline SSPs that undergo solid-state transformation at different temperatures and rates.6 Under identical conditions, these different SSPs yield thin films with distinct grain size, surface roughness, and pinhole density. This can be attributed to the complex interplay between the kinetics of nucleation, solvent evaporation, grain growth, coarsening, and decomposition.6 Solvent chemistry is another important parameter that has more recently been explored to improve solution-processed perovskite films. Use of coordinating solvents adds additional complexity, but also the opportunity to exploit solvent−solute interactions. DMSO ((CH3)2SO) is well-known as a coordinating solvent, both as a monodentate ligand through either atom of the S−O bond and through hydrogen bonding,22 and will serve as a model system for discussion. Other common solvents used in perovskite thin film preparation, such as DMF and γbutyrolactone (GBL), can also act as coordinating ligands that influence solution chemistry and subsequent thin film quality and morphology.3,7,8 In the solid-state compound PbI2(DMSO)2, DMSO molecules are intercalated between 2-dimensional PbI2 inorganic sheets.23 Incorporation of the organic molecules expands the PbI2 lattice as indicated by X-ray reflections at lower angles (2θ ∼ 10°, Figure 4a) relative to neat PbI2. This culminates in an effective “preloading” of the PbI2 lattice, which is then more C

DOI: 10.1021/acs.accounts.5b00455 Acc. Chem. Res. XXXX, XXX, XXX−XXX

Article

Accounts of Chemical Research

amenable to intramolecular exchange reactions or processes (IEPs) as illustrated schematically in Figure 4b. The expanded lattice of PbI2-DMSO composites yields thicker films compared to neat PbI2 under the same processing conditions. As-deposited films can then be rapidly converted (on the order of 1−2 min) to hybrid perovskites with minimal volume expansion by contact with an isopropanolic solution of organohalide precursor (Figure 4c). In order to achieve this rapid conversion, PbI2 and DMSO must be in an equimolar ratio, otherwise annealing above 100 °C is required to drive off excess DMSO.9,10,24 Intramolecular exchange processing has significantly improved perovskite solar cell performance over traditional fabrication routes (Figure 4d), yielding certified efficiency over 20%.10 Similar solvent-precursor manipulation strategies have been extended to films prepared by one step deposition, where processing temperatures and times are significantly reduced (65 °C, 1 min) due to the volatility of DMSO molecules in the 1:1 adduct structure.9 We emphasize that tuning solvent chemistry is still relatively uncharted with regard to hybrid perovskite synthesis. We anticipate that additional contributions in this area will further improve device performance and reproducibility. Inverse Temperature Crystallization of Perovskite Single Crystals

Figure 3. (a) Powder XRD diffraction patterns of 3:1 and 1:1 mpAl2O3/CH3NH3PbI3 films at various stages of formation. PbI2 and CH3NH3I precursors are shown for reference. Black and red vertical lines indicate prominent reflections from crystalline solid-state precursor (SSP) and CH3NH3PbI3 phases, respectively. A narrower 2θ range is given to the right for clarity. Scanning electron micrographs of annealed (b) 1:1 and (c) 3:1 CH3NH3PbI3 films on FTO. Adapted with permission from ref 3. Copyright 2015 American Chemical Society.

Solution-grown hybrid perovskite single crystals are comparable in transport properties and defect densities to high-end optoelectronic-grade inorganic semiconductors.25 Beyond their properties, single crystals are ideal model systems to investigate the role of plumbate complexes on crystallization mechanisms. Their highly ordered structure and comparably slower growth affords a clearer view of the crystallization processes and final products that are otherwise challenging to elucidate in thin films,

Figure 4. (a) XRD patterns of PbI2(DMSO)2 and vacuum-annealed PbI2(DMSO) powders and an as-deposited thin film on fused quartz using PbI2(DMSO). (b) Illustration of IEP within the PbI2 framework involving substitution of DMSO molecules with formamidinium iodide. (c) Scanning electron micrographs of PbI2 (conventional) and PbI2(DMSO) (IEP) thin films before and after conversion to perovskite. (d) J−V curves of perovskite solar cells prepared by IEP and conventional sequential deposition. From ref 10. Reprinted with permission from AAAS. D

DOI: 10.1021/acs.accounts.5b00455 Acc. Chem. Res. XXXX, XXX, XXX−XXX

Article

Accounts of Chemical Research

crystallization (ITC) (Figure 5b).29 With ITC, hybrid perovskite single crystals grow at an astonishing rate of ∼20 mm3/h, an order of magnitude faster than traditional techniques. These bulk crystals, grown in hours (Figure 5c and d), exhibit carrier transport properties and defect trap-state densities comparable to those grown over weeks by classical methods (Table 1).

which are riddled with defects and microstructure-level inhomogeneities.26,27 Perovskite single crystals are typically grown by loss of solubility upon diffusion of an antisolvent into solution25 or upon cooling of a saturated solution.28 Unusual loss of solubility upon heating, known as retrograde solubility, was recently observed in hybrid perovskites.29 The coordination of solvent molecules and plumbates (vide supra) results in this rarely observed phenomenon, which is only displayed by a few other materials. The paradoxical retrograde solubility of hybrid perovskites is strongly influenced by the solvent. For instance, CH3NH3PbCl3 in DMSO or mixtures of DMSO/DMF show the remarkable loss of solubility by heating the solution (Figure 5a).30 On the other

Table 1. Charge Carrier Mobility and Trap-State Density Estimated by Space Charge Limited Current Measurements in CH3NH3PbX3 Single Crystals Grown by Different Techniques growth technique

charge carrier mobility (cm2 V−1 s−1)

trap-state density (cm−3)

67.2 ± 7.3

1.4 × 1010

38

5.8 × 109

164 ± 25

3.6 × 1010

inverse temperature crystallization29 antisolvent vapor-assisted crystallization25 top-seeded solution growth32

Designing high-quality single crystal growth by ITC requires a careful analysis of the temperature dependent solubility of complexes in various solvent and solvent mixtures to maximize both yield and control over growth kinetics. Retrograde solubility and ITC are prime examples demonstrating the new modes of crystallization that could be achieved by taking advantage of the rich variety of plumbate complexes with solvent molecules.



INSTABILITY OF ORGANIC−INORGANIC PEROVSKITES The critical role that humidity plays in the degradation of perovskite solar cell performance was realized early on in their development.33 The generally accepted overall reaction for CH3NH3PbI3 degradation is shown in reaction 3.12

Figure 5. (a) Solubility data for CH3NH3PbCl3 powder in DMSO− DMF (1:1 v/v) at different temperatures. Inset−transparent single crystal grown from CH3NH3PbCl3 solution in DMSO−DMF (1:1 v/v). Adapted with permission from ref 30. Copyright 2015 American Chemical Society. (b) Schematic representation of the ITC apparatus. The solution is heated from room temperature and kept at an elevated temperature (80 °C for CH3NH3PbBr3 and 110 °C for CH3NH3PbI3) to induce crystallization.29 (c,d) CH3NH3PbI3 and CH3NH3PbBr3 crystal growth at different time intervals.29

CH3NH3PbI3(s) → CH3NH 2(g) + HI(g) + PbI 2(s)

(3)

In this mechanism, H2O deprotonates CH3NH3+ which results in the formation of the subsequent degradation products. This results in a widening of the band gap and change in surface morphology (Figure 6a and b). While the basic outline of the degradation pathway is quite clear, it is only more recently that the full complexity of the initial H2O−CH3NH3PbI3 interaction has begun to be unveiled. For instance, degradation of CH3NH3PbI3 under ambient conditions was recently shown to result in compounds other than PbI2, such as PbCO3 and αPbO.34 Moreover, the rate of formation of oxide and carbonate degradation products in neat PbI2 is different than CH3NH3PbI3, suggesting compounds other than stoichiometric PbI2 may serve as decomposition intermediates under certain conditions. CH3NH3PbI3 has been shown to complex with H2O, forming both monohydrate CH 3 NH 3 PbI 3 ·H 2 O 16 and dihydrate (CH3NH3)4PbI6·2H2O.35 Recent crystallographic studies of moisture exposed CH3NH3PbI3 films have revealed the formation of these hydrates as an integral part of the degradation mechanism (Figure 6c−f), outlined in reaction 4.36,37

hand, CH3NH3PbBr3 and CH3NH3PbI3 demonstrate retrograde solubility in DMF and GBL, respectively.31 Noteworthy, perovskite precursors individually do not exhibit any retrograde solubility behavior, indicating that the plumbate complexes are necessary to induce this phenomenon. Since the presence of a specific solvent is also necessary to induce retrograde solubility, the precise nature of lead complexes in the precursor solution dictates the crystallization process. Complexes formed between plumbates and DMSO retard the crystallization of perovskite.24 In such cases, amorphous films crystallize into the perovskite phase only after being annealed at temperatures sufficient to evaporate coordinated solvent molecules. Similarly, the dissociation of these complexes in solution is a pathway to form single crystals. Consequently, a solvent that forms stable complexes with plumbates and solvent molecules, which in turn dissociate at elevated temperatures (in solution), will commence retrograde solubility. For example, in the case of CH3NH3PbI3, retrograde solubility occurs only in GBL, indicating that GBL-based complexes dissociate at elevated temperatures whereas DMF- or DMSO-based complexes seem to be more stable and may not dissociate without evaporating the solvent. Overall, the crystallization reaction is endothermic, with molecules moving from a complex of higher binding energy to a crystalline precipitate state of lower cohesive energy. Retrograde solubility of hybrid perovskites enables a novel nonclassical pathway to crystallization: inverse temperature

4CH3NH3PbI3 + 4H 2O ↔ 4[CH3NH3PbI3·H 2O] ↔ (CH3NH3)4 PbI6 ·2H 2O + 3PbI 2 + 2H 2O

(4)

Moisture-exposed CH3NH3PbI3 single crystals also undergo a hydration reaction. However, this process is significantly retarded in single crystals compared to that in thin films, for which their abundant grain boundaries enable a faster and a E

DOI: 10.1021/acs.accounts.5b00455 Acc. Chem. Res. XXXX, XXX, XXX−XXX

Article

Accounts of Chemical Research

Figure 6. (a) UV−visible absorption spectra of CH3NH3PbI3 films stored in the dark at room temperature in 90% relative humidity and a pristine PbI2 film. (b) Photographs and SEM images of pristine (left) and hydrated (right) films. Adapted with permission from ref 36. Copyright 2015 American Chemical Society. (c) Identification of the composition of the hydrate species grown on CH3NH3PbI3 single crystals (at long exposure to water vapors) and of CH3NH3PbI3 hydrate crystals (polycrystalline, obtained from solution) by X-ray diffraction (patterns in black and gray). (d−f) Structure of CH3NH3PbI3, CH3NH3PbI3·H2O, and (CH3NH3)4PbI6·2H2O, respectively. Adapted with permission from ref 37. Copyright 2015 American Chemical Society.

higher degree of penetration of water molecules.37 The hydrogen bonding interactions between the organic cation and halide ions of the perovskite lattice provide structural stability;17 however, calculations have shown that the polar water molecules readily intercalate into the perovskite lattice because of their ability to form strong hydrogen bonds to the lattice iodides and, to a lesser extent, the methylammonium cations.38 Furthermore, it is proposed that halide-to-metal charge transfer undergone upon light absorption lessens the charge density at the X sites, thereby weakening the CH3NH3···X hydrogen bonding interactions and causing an increased rate of degradation under illumination in a humid atmosphere.15,36 The observed decrease in photovoltaic performance which results from moisture exposure is hypothesized to result, at least initially, from inhibited charge collection caused by formation of insulating hydrates and grain boundaries in the perovskite film.36,37 Intriguingly, calculations have shown that a lower hydrate phase (4:1 CH3NH3PbI3/H2O) may also form, leading to a 50 meV increase in band gap and ∼1% volume expansion.39 Targeted experimental studies are required to elucidate the existence of this hydrate phase and its potential modulation of perovskite optoelectronic properties. Current evidence points to the facile reversibility of CH 3 NH 3 PbI 3 hydration. 36,37 Restoration of the neat CH3NH3PbI3 can occur by removal of ambient moisture. This has been demonstrated in working photovoltaic devices37 and illuminates the apparent discrepancy recently arising in the literature as to the role of humidity in solar cell fabrication.40−43

These seemingly contradictory claims can be harmonized by the reversibility of hydration. As films are cast using certain techniques and precursor chemistries, water may be beneficial by complexing with the precursors, both in solution and in the solid state. For example, average grain size was shown to increase from ∼200 nm to >500 nm by annealing thin films prepared by single-step deposition in ambient conditions (35% relative humidity) instead of inert atmosphere.44 Conversely, moisture can inhibit film formation when using other techniques such as two-step sequential deposition of PbI 2 and CH 3 NH 3 I precursors.45 The wide range of perovskite solar cell processing conditions makes overly general conclusions about moisture during fabrication somewhat moot. Nevertheless, development of new processing techniques and/or precursor chemistries which are humidity agnostic and result in highly efficient devices appears possible due to the reversibility of CH3NH3PbI3 hydration.42 Degradation of the organic component within hybrid perovskites occurs at relatively low temperatures (∼140 °C for CH3NH3PbI3), which may impact long-term operational stability, especially in hot climates. With methylammonium cations, sublimation of CH3NH2 and HI eventually leads to formation of PbI2.46 In contrast, CsPbI3 is typically processed at 250 °C, highlighting the increased thermal stability of allinorganic perovskites.47 The ability to carry out solid-state exchange of halide ions further highlights the susceptibility of hybrid perovskite to F

DOI: 10.1021/acs.accounts.5b00455 Acc. Chem. Res. XXXX, XXX, XXX−XXX

Article

Accounts of Chemical Research undergo chemical change after their initial formation. For example, exposure of CH3NH3PbI3 to Br− ions causes rapid transformation to CH3NH3PbBr3.48 An upside of this facile exchange is the ability to tune the absorption and emission properties across the entire visible region.49 As discussed earlier, the principles of organometallic chemistry dictate the halide exchange in these systems.

perovskite solar cells, and the application of those devices for electricity and solar fuels generation. Makhsud I. Saidaminov was born in Tajikistan in 1988. He received his Ph.D. in 2013 from Lomonosov Moscow State University, Russia. He is currently a Postdoctoral Fellow in the Physical Sciences and Engineering division at KAUST, Saudi Arabia. His research focuses on the engineering of materials, particularly hybrid perovskite single crystals, and their integration into optoelectronic devices.



FUTURE PROSPECTS As the field of hybrid perovskites continues to grow, it is of utmost importance that studies continue to focus on molecular scale interactions that dictate thin film and single crystal formation mechanics. It is also imperative to fully understand how microscopic characteristics (e.g., dynamic motion of organic cations, morphology/phase, ion migration) influence macroscopic properties through combined experimental and computational approaches. Addressing the downside of low formation energies, namely, the relative ease with which hybrid perovskites like CH3NH3PbI3 break down, is a top priority moving forward. The question as to whether CH3NH3PbI3 and its analogues can be engineered to meet stringent international standards for outdoor photovoltaic operations may be one of simple thermodynamics. Recent calculations have shown that the disproportionation reaction CH3NH3PbI3 → CH3NH3I + PbI2 is exothermic.50 Issues of device longevity can be addressed in two ways. The first method is an external approach, where engineering of interfacial contacts and the device stack in general lends stability to perovskite active layers.51 The second, more arduous method is an internal approach that involves modifying the underlying chemistry to improve intrinsic stability. Incorporation of polar neutral molecules like water and DMF into the perovskite lattice is partially stabilized by hydrogen bonding between the organic cation and guest molecules.16 Lessening this interaction can improve the stability in the presence of moisture. The use of aprotic cations such as (CH3)4N+ may impede moisture-induced degradation,12 but size constraints limit the number of such compounds that are compatible with the 3-dimensional perovskite structure. Mixed-halide perovskites (e.g., CH3NH3PbI3−xBrx) exhibit reduced chemical lability with minimal impact on optoelectronic properties.33,50 However, such alloys are known to dissociate when irradiated.52 It may be that abandoning the organic component altogether is the most viable solution. Regardless of how instability is addressed, establishing design rules for hybrid perovskites through study of formation and degradation mechanisms can help direct synthesis of more resilient compounds, ideally striking a balance between low-cost processing, long-term durability, and exceptional device performance.



Jeffrey A. Christians was born in Grand Rapids, Michigan, in 1988. He received his B.S.E. in 2010 from Calvin College and his Ph.D. from the University of Notre Dame in 2015. He is currently a postdoctoral researcher in the Chemistry & Nanoscience Center at the National Renewable Energy Laboratory in Golden, CO. Here his research focuses on understanding the relationship between the structure and key optoelectronic properties of hybrid organic−inorganic perovskites. Osman M. Bakr was born in Saudi Arabia in 1981. He holds a B.Sc. in Materials Science and Engineering from MIT (2003) as well as a M.S. and Ph.D. in Applied Physics from Harvard University (2009). He is currently an Associate Professor of Materials Science and Engineering, SABIC Presidential Career Development Chair, at KAUST, Saudi Arabia. His research group focuses on the study of hybrid organic− inorganic nanoparticles and materials; particularly advancing their synthesis and self-assembly for applications in photovoltaics and optoelectronics. Prashant V. Kamat is the Rev. John A. Zahm, C.S.C., Professor of Science at the University of Notre Dame. He earned his Ph.D. in Physical Chemistry from Bombay University in 1979 and was a postdoctoral researcher at Boston University and the University of Texas at Austin before joining Notre Dame.



ACKNOWLEDGMENTS We acknowledge the support of King Abdullah University of Science and Technology (KAUST) through the Award OCRF2014-CRG3-2268. The previously published material discussed in this Account was supported by the U.S. Department of Energy Office of Science, Office of Basic Energy Sciences under Award Number DE-FC02-04ER15533. This is contribution number NDRL 5090 from the Notre Dame Radiation Laboratory.



REFERENCES

(1) Schlom, D. G.; Chen, L. Q.; Pan, X.; Schmehl, A.; Zurbuchen, M. A. A Thin Film Approach to Engineering Functionality into Oxides. J. Am. Ceram. Soc. 2008, 91, 2429−2454. (2) Reller, A.; Williams, T. Perovskites - Chemical Chameleons. Chem. Br. 1989, 25, 1227−1230. (3) Manser, J. S.; Reid, B.; Kamat, P. V. Evolution of Organic− Inorganic Lead Halide Perovskite from Solid-State Iodoplumbate Complexes. J. Phys. Chem. C 2015, 119, 17065−17073. (4) Stamplecoskie, K. G.; Manser, J. S.; Kamat, P. V. Dual Nature of the Excited State in Organic−Inorganic Lead Halide Perovskites. Energy Environ. Sci. 2015, 8, 208−215. (5) Williams, S. T.; Zuo, F.; Chueh, C.-C.; Liao, C.-Y.; Liang, P.-W.; Jen, A. K.-Y. Role of Chloride in the Morphological Evolution of Organo-Lead Halide Perovskite Thin Films. ACS Nano 2014, 8, 10640− 10654. (6) Moore, D. T.; Sai, H.; Tan, K. W.; Smilgies, D.-M.; Zhang, W.; Snaith, H. J.; Wiesner, U.; Estroff, L. A. Crystallization Kinetics of Organic-Inorganic Trihalide Perovskites and the Role of the Lead Anion in Crystal Growth. J. Am. Chem. Soc. 2015, 137, 2350−2358. (7) Yan, K.; Long, M.; Zhang, T.; Wei, Z.; Chen, H.; Yang, S.; Xu, J. Hybrid Halide Perovskite Solar Cell Precursors: The Colloidal Chemistry and Coordination Engineering behind Device Processing for High Efficiency. J. Am. Chem. Soc. 2015, 137, 4460−4468.

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Notes

The authors declare no competing financial interest. Biographies Joseph S. Manser was born in Youngstown, Ohio, in 1989. He received his B.Sc. in 2011 from Catawba College and is currently pursuing a Ph.D. at the University of Notre Dame. His research focuses on the fundamental photophysics and structure−property relationships that underlie operation of next-generation optoelectronic devices, including G

DOI: 10.1021/acs.accounts.5b00455 Acc. Chem. Res. XXXX, XXX, XXX−XXX

Article

Accounts of Chemical Research (8) Guo, Y.; Shoyama, K.; Sato, W.; Matsuo, Y.; Inoue, K.; Harano, K.; Liu, C.; Tanaka, H.; Nakamura, E. Chemical Pathways Connecting Lead(II) Iodide and Perovskite via Polymeric Plumbate(II) Fiber. J. Am. Chem. Soc. 2015, 137, 15907−15914. (9) Ahn, N.; Son, D.-Y.; Jang, I.-H.; Kang, S. M.; Choi, M.; Park, N.-G. Highly Reproducible Perovskite Solar Cells with Average Efficiency of 18.3% and Best Efficiency of 19.7% Fabricated via Lewis Base Adduct of Lead(II) Iodide. J. Am. Chem. Soc. 2015, 137, 8696−8699. (10) Yang, W. S.; Noh, J. H.; Jeon, N. J.; Kim, Y. C.; Ryu, S.; Seo, J.; Seok, S. I. High-Performance Photovoltaic Perovskite Layers Fabricated through Intramolecular Exchange. Science 2015, 348, 1234−1237. (11) Leguy, A. M. a.; Frost, J. M.; McMahon, A. P.; Sakai, V. G.; Kochelmann, W.; Law, C.; Li, X.; Foglia, F.; Walsh, A.; O’Regan, B. C.; Nelson, J.; Cabral, J. T.; Barnes, P. R. F. The Dynamics of Methylammonium Ions in Hybrid Organic−inorganic Perovskite Solar Cells. Nat. Commun. 2015, 6, 7124. (12) Frost, J. M.; Butler, K. T.; Brivio, F.; Hendon, C. H.; van Schilfgaarde, M.; Walsh, A. Atomistic Origins of High-Performance in Hybrid Halide Perovskite Solar Cells. Nano Lett. 2014, 14, 2584−2590. (13) Even, J.; Pedesseau, L.; Katan, C. Analysis of Multivalley and Multibandgap Absorption and Enhancement of Free Carriers Related to Exciton Screening in Hybrid Perovskites. J. Phys. Chem. C 2014, 118, 11566−11572. (14) Liu, S.; Zheng, F.; Koocher, N. Z.; Takenaka, H.; Wang, F.; Rappe, A. M. Ferroelectric Domain Wall Induced Band Gap Reduction and Charge Separation in Organometal Halide Perovskites. J. Phys. Chem. Lett. 2015, 6, 693−699. (15) Gottesman, R.; Haltzi, E.; Gouda, L.; Tirosh, S.; Bouhadana, Y.; Zaban, A.; Mosconi, E.; De Angelis, F. Extremely Slow Photoconductivity Response of CH3NH3PbI3 Perovskites Suggesting Structural Changes under Working Conditions. J. Phys. Chem. Lett. 2014, 5, 2662−2669. (16) Hao, F.; Stoumpos, C. C.; Liu, Z.; Chang, R. P. H.; Kanatzidis, M. G. Controllable Perovskite Crystallization at a Gas−Solid Interface for Hole Conductor-Free Solar Cells with Steady Power Conversion Efficiency over 10%. J. Am. Chem. Soc. 2014, 136, 16411−16419. (17) Quarti, C.; Grancini, G.; Mosconi, E.; Bruno, P.; Ball, J. M.; Lee, M. M.; Snaith, H. J.; Petrozza, A.; Angelis, F. De. The Raman Spectrum of the CH3NH3PbI3 Hybrid Perovskite: Interplay of Theory and Experiment. J. Phys. Chem. Lett. 2014, 5, 279−284. (18) Horváth, O.; Mikó, I. Spectra, Equilibrium and Photoredox Chemistry of Tri- and Tetraiodoplumbate(II) Complexes in Acetonitrile. J. Photochem. Photobiol., A 1998, 114, 95−101. (19) Grancini, G.; Marras, S.; Prato, M.; Giannini, C.; Quarti, C.; De Angelis, F.; De Bastiani, M.; Eperon, G. E.; Snaith, H. J.; Manna, L.; Petrozza, A. The Impact of the Crystallization Processes on the Structural and Optical Properties of Hybrid Perovskite Films for Photovoltaics. J. Phys. Chem. Lett. 2014, 5, 3836−3842. (20) Manser, J. S.; Kamat, P. V. Band Filling with Free Charge Carriers in Organometal Halide Perovskites. Nat. Photonics 2014, 8, 737−743. (21) Yang, S.; Zheng, Y. C.; Hou, Y.; Chen, X.; Chen, Y.; Wang, Y.; Zhao, H.; Yang, H. G. Formation Mechanism of Freestanding CH3 NH3PbI3 Functional Crystals: In Situ Transformation vs Dissolution−Crystallization. Chem. Mater. 2014, 26, 6705−6710. (22) Reynolds, W. L. Dimethyl Sulfoxide in Inorganic Chemistry. In Progress in Inorganic Chemistry; Lippard, S., Ed.; John Wiley & Sons, Inc.: New York, 1970; pp 1−100. (23) Miyamae, H.; Numahata, Y.; Nagata, M. The Crystal Structure of Lead(II) Iodide-Dimethylsulfoxide(1/2), PbI2(dmso)2. Chem. Lett. 1980, 9, 663−664. (24) Jeon, N. J.; Noh, J. H.; Kim, Y. C.; Yang, W. S.; Ryu, S.; Seok, S., Il. Solvent Engineering for High-Performance Inorganic-Organic Hybrid Perovskite Solar Cells. Nat. Mater. 2014, 13, 897−903. (25) Shi, D.; Adinolfi, V.; Comin, R.; Yuan, M.; Alarousu, E.; Buin, A.; Chen, Y.; Hoogland, S.; Rothenberger, A.; Katsiev, K.; Losovyj, Y.; Zhang, X.; Dowben, P. A.; Mohammed, O. F.; Sargent, E. H.; Bakr, O. M. Low Trap-State Density and Long Carrier Diffusion in Organolead Trihalide Perovskite Single Crystals. Science 2015, 347, 519−522.

(26) de Quilettes, D. W.; Vorpahl, S. M.; Stranks, S. D.; Nagaoka, H.; Eperon, G. E.; Ziffer, M. E.; Snaith, H. J.; Ginger, D. S. Impact of Microstructure on Local Carrier Lifetime in Perovskite Solar Cells. Science 2015, 348, 683−686. (27) Bischak, C. G.; Sanehira, E. M.; Precht, J. T.; Luther, J. M.; Ginsberg, N. S. Heterogeneous Charge Carrier Dynamics in Organic− Inorganic Hybrid Materials: Nanoscale Lateral and Depth-Dependent Variation of Recombination Rates in Methylammonium Lead Halide Perovskite Thin Films. Nano Lett. 2015, 15, 4799−4807. (28) Dang, Y.; Liu, Y.; Sun, Y.; Yuan, D.; Liu, X.; Lu, W.; Liu, G.; Xia, H.; Tao, X. Bulk Crystal Growth of Hybrid Perovskite Material CH3NH3PbI3. CrystEngComm 2015, 17, 665−670. (29) Saidaminov, M. I.; Abdelhady, A. L.; Murali, B.; Alarousu, E.; Burlakov, V. M.; Peng, W.; Dursun, I.; Wang, L.; He, Y.; Maculan, G.; Goriely, A.; Wu, T.; Mohammed, O. F.; Bakr, O. M. High-Quality Bulk Hybrid Perovskite Single Crystals within Minutes by Inverse Temperature Crystallization. Nat. Commun. 2015, 6, 7586. (30) Maculan, G.; Sheikh, A. D.; Abdelhady, A. L.; Saidaminov, M. I.; Haque, M. A.; Murali, B.; Alarousu, E.; Mohammed, O. F.; Wu, T.; Bakr, O. M. CH3NH3PbCl3 Single Crystals: Invese Temperature Crystallization and Visible-Blind UV-Photodetector. J. Phys. Chem. Lett. 2015, 6, 3781−3786. (31) Saidaminov, M. I.; Abdelhady, A. L.; Maculan, G.; Bakr, O. M. Retrograde Solubility of Formamidinium and Methylammonium Lead Halide Perovskites Enabling Rapid Single Crystal Growth. Chem. Commun. 2015, 51, 17658−17661. (32) Dong, Q.; Fang, Y.; Shao, Y.; Mulligan, P.; Qiu, J.; Cao, L.; Huang, J. Electron-Hole Diffusion Lengths > 175 μm in Solution-Grown CH3NH3PbI3 Single Crystals. Science 2015, 347, 967−970. (33) Noh, J. H.; Im, S. H.; Heo, J. H.; Mandal, T. N.; Seok, S., Il. Chemical Management for Colorful, Efficient, and Stable InorganicOrganic Hybrid Nanostructured Solar Cells. Nano Lett. 2013, 13, 1764− 1769. (34) Huang, W.; Manser, J. S.; Kamat, P. V.; Ptasinska, S. Evolution of Chemical Composition, Morphology, and Photovoltaic Efficiency of CH3NH3PbI3 Perovskite under Ambient Conditions. Chem. Mater. 2016, 28, 303−311. (35) Vincent, B. R.; Robertson, K. N.; Cameron, T. S.; Knop, O. Alkylammonium Lead Halides. Part 1. Isolated PbI64− Ions in (CH3NH3)4PbI6•2H2O. Can. J. Chem. 1987, 65, 1042−1046. (36) Christians, J. A.; Miranda Herrera, P. A.; Kamat, P. V. Transformation of the Excited State and Photovoltaic Efficiency of CH3NH3PbI3 Perovskite upon Controlled Exposure to Humidified Air. J. Am. Chem. Soc. 2015, 137, 1530−1538. (37) Leguy, A.; Hu, Y.; Campoy-Quiles, M.; Alonso, M. I.; Weber, O. J.; Azarhoosh, P.; van Schilfgaarde, M.; Weller, M. T.; Bein, T.; Nelson, J.; Docampo, P.; Barnes, P. R. F. The Reversible Hydration of CH3NH3PbI3 in Films, Single Crystals and Solar Cells. Chem. Mater. 2015, 27, 3397−3407. (38) Zhang, L.; Sit, P. H.-L. Ab Initio Study of Interaction of Water, Hydroxyl Radicals, and Hydroxide Ions with CH3NH3PbI3 and CH3NH3PbBr3 Surfaces. J. Phys. Chem. C 2015, 119, 22370−22378. (39) Mosconi, E.; Azpiroz, J. M.; De Angelis, F. Ab Initio Molecular Dynamics Simulations of Methylammonium Lead Iodide Perovskite Degradation by Water. Chem. Mater. 2015, 27, 4885−4892. (40) Bass, K. K.; McAnally, R. E.; Zhou, S.; Djurovich, P. I.; Thompson, M.; Melot, B. Influence of Moisture on the Preparation, Crystal Structure, and Photophysical Properties of Organohalide Perovskites. Chem. Commun. 2014, 50, 15819−15822. (41) Zhou, H.; Chen, Q.; Li, G.; Luo, S.; Song, T. -b.; Duan, H.-S.; Hong, Z.; You, J.; Liu, Y.; Yang, Y. Interface Engineering of Highly Efficient Perovskite Solar Cells. Science 2014, 345, 542−546. (42) Conings, B.; Babayigit, A.; Vangerven, T.; D’Haen, J.; Manca, J.; Boyen, H.-G. The Impact of Precursor Water Content on SolutionProcessed Organometal Halide Perovskite Films and Solar Cells. J. Mater. Chem. A 2015, 3, 19123−19128. (43) Wozny, S.; Yang, M.; Nardes, A. M.; Mercado, C. C.; Ferrere, S.; Reese, M. O.; Zhou, W.; Zhu, K. Controlled Humidity Study on the H

DOI: 10.1021/acs.accounts.5b00455 Acc. Chem. Res. XXXX, XXX, XXX−XXX

Article

Accounts of Chemical Research Formation of Higher Efficiency Formamidinium Lead Triiodide-Based Solar Cells. Chem. Mater. 2015, 27, 4814−4820. (44) You, J.; Yang, Y.; Hong, Z.; Song, T.-B.; Meng, L.; Liu, Y.; Jiang, C.; Zhou, H.; Chang, W.-H.; Li, G.; Yang, Y. Moisture Assisted Perovskite Film Growth for High Performance Solar Cells. Appl. Phys. Lett. 2014, 105, 183902. (45) Ko, H.-S.; Lee, J.-W.; Park, N.-G. 15.76% Efficiency Perovskite Solar Cells Prepared under High Relative Humidity: Importance of PbI2 Morphology in Two-Step Deposition of CH3NH3PbI3. J. Mater. Chem. A 2015, 3, 8808−8815. (46) Philippe, B.; Park, B.; Lindblad, R.; Oscarsson, J.; Ahmadi, S.; Johansson, E. M. J.; Rensmo, H. Chemical and Electronic Structure Characterization of Lead Halide Perovskites and Stability Behavior under Different ExposuresA Photoelectron Spectroscopy Investigation. Chem. Mater. 2015, 27, 1720−1731. (47) Kulbak, M.; Cahen, D.; Hodes, G. How Important Is the Organic Part of Lead Halide Perovskite Photovoltaic Cells? Efficient CsPbBr3 Cells. J. Phys. Chem. Lett. 2015, 6, 2452−2456. (48) Solis-Ibarra, D.; Smith, I. C.; Karunadasa, H. I. Post-Synthetic Halide Conversion and Selective Halogen Capture in Hybrid Perovskites. Chem. Sci. 2015, 6, 4054−4059. (49) Akkerman, Q. a.; D’Innocenzo, V.; Accornero, S.; Scarpellini, A.; Petrozza, A.; Prato, M.; Manna, L. Tuning the Optical Properties of Cesium Lead Halide Perovskite Nanocrystals by Anion Exchange Reactions. J. Am. Chem. Soc. 2015, 137, 10276−10281. (50) Zhang, Y.-Y.; Chen, S.; Xu, P.; Xiang, H.; Gong, X.-G.; Walsh, A.; Wei, S. Intrinsic Instability of the Hybrid Halide Perovskite Semiconductor CH3NH3PbI3. 2015, arXiv:1506.01301 [cond-mat.mtrl-sci]. arXiv.org e-Print archive. http://arxiv.org/abs/1506.01301. (51) Mei, A.; Li, X.; Liu, L.; Ku, Z.; Liu, T.; Rong, Y.; Xu, M.; Hu, M.; Chen, J.; Yang, Y.; Grätzel, M.; Han, H. A Hole-Conductor−free, Fully Printable Mesoscopic Perovskite Solar Cell with High Stability. Science 2014, 345, 295−298. (52) Hoke, E. T.; Slotcavage, D. J.; Dohner, E. R.; Bowring, A. R.; Karunadasa, H. I.; McGehee, M. D. Reversible Photo-Induced Trap Formation in Mixed-Halide Hybrid Perovskites for Photovoltaics. Chem. Sci. 2015, 6, 613−617.

I

DOI: 10.1021/acs.accounts.5b00455 Acc. Chem. Res. XXXX, XXX, XXX−XXX